Anda di halaman 1dari 37

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)
P1: IBC
10.1146/annurev.fluid.35.101101.161144

Annu. Rev. Fluid Mech. 2003. 35:13567


doi: 10.1146/annurev.fluid.35.101101.161144
c 2003 by Annual Reviews. All rights reserved
Copyright

MIXING EFFICIENCY IN STRATIFIED SHEAR FLOWS


W. R. Peltier
Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org
by University of Toronto on 06/14/06. For personal use only.

Department of Physics, University of Toronto, Toronto, Ontario M5S 1A7 Canada;


e-mail: peltier@atmosp.physics.utoronto.ca

C. P. Caulfield
Department of Mechanical and Aerospace Engineering, University of California,
San Diego, La Jolla, California 92093-0411; e-mail: cpc@mae.ucsd.edu

Key Words stratified flows, transition


Abstract The issue of the physical mechanism(s) that control the efficiency with
which the density field in stably stratified fluid is mixed by turbulent processes has remained enigmatic. Similarly enigmatic has been an explanation of the numerical value
of 0.2, which is observed to characterize this efficiency experimentally. We review
recent work on the turbulence transition in stratified parallel flows that demonstrates
that this value is not only numerically predictable but also that it is expected to be a
nonmonotonic function of the Richardson number that characterizes preturbulent stratification strength. This value of the mixing efficiency appears to be characteristic of the
late-time behavior of the turbulent flow that develops after an initially laminar shear
flow has undergone the transition to turbulence through an intermediate instability of
Kelvin-Helmholtz type.

1. INTRODUCTION
In comparison with the relatively simple problem of thermal convection, the problem of understanding the mechanics of the turbulence transition in an initially laminar free shear layer involves a considerable increase in complexity. Whereas
in the problem of thermal convection the strength of the thermal forcing is most
often kept fixed by fixing the appropriate Rayleigh number and then measuring the
corresponding Nusselt number (nondimensional heat transfer) in a statistical equilibrium state, in the case of the free shear layer, mixing is forced to occur through a
highly time-dependent sequence of specific events that originate through a simple
mechanically driven instability whose strength depends not only on the magnitude
but also the structure of the vertical shear in horizontal velocity. In this problem
the correct nondimensional measure of the strength of the initial mechanical forcing is the so-called gradient Richardson number Ri(z), which is defined as
Ri(z) =
0066-4189/03/0115-0135$14.00

(gd`og(z)/dz)

N2
=
,
2
(dV/dz)
(dV/dz)2

(1)
135

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

136

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

in which N2(z) is the square of the Brunt-Vaisala frequency, (z)

is the height
variation of the density field in the initially parallel-flow basic state, and V(z) is
the initial height variation of background horizontal velocity. The important theorem of Miles (1961) and Howard (1961) assures us that this stratified parallel
flow will be stable against small amplitude fluctuations so long as Ri(z) > 0.25
at all heights. However, as soon as Ri(z) < 0.25 at some level in the parallel
flow, a rich variety of instabilities may occur (see Drazin & Reid 1981). Although
circumstances in which the stratified shear instability that arises when the MilesHoward criterion for stability is violated may produce coherent structures other
than the vortical disturbances of Kelvin and Helmholtz (KH) (see Drazin & Reid
1981), the development of these so-called KH instabilities appears to constitute
the usual first step toward the onset of vigorous mixing of the density field, and
they have been observed both in the atmosphere and oceans (see, e.g., Woods
1968, De Silva et al. 1996). KH instabilities are the inevitable consequence of
instability in circumstances in which either the stratification is sufficiently weak
or the thickness of the layer of strong, inflectional shear (see Drazin & Reid
1981) is similar to the thickness of any coincident concentration in the (stable)
density gradient. Only when the latter thickness is significantly smaller than the
former, and the stratification is sufficiently strong, does the characteristic instability bifurcate into the more exotic Holmboe instability (Holmboe 1962, Browand
& Winant 1973, Koop & Browand 1979, Smyth et al. 1988, Lawrence et al.
1991).
The Holmboe instability is characterized at finite amplitude (see Smyth et al.
1988) by vortical disturbances displaced above and below the region of strong
density gradient that propagate in opposite directions relative to the mean flow
unlike the KH instability, which is stationary in a frame of reference moving with
the mean flow. The Holmboe instability also induces significant overturning of
the region of strong density gradient. Experimental evidence (Browand & Winant
1973, Koop & Browand 1979, Strang & Fernando 2001) suggests that flows that
exhibit Holmboe instabilities alone cause relatively weak mixing compared to the
intense mixing associated with high Reynolds number and low Richardson number
KH instabilities. However, some recent numerical evidence (Smyth & Winters
2002) suggests that the long time evolution of the instability (in the absence of
transition) may nevertheless induce significant mixing, and experimental evidence
(see Strang & Fernando 2001) suggests that the Holmboe instability may play a
significant role in enhancing the mixing associated with KH breakdown at higher
Richardson numbers.
Because KH instabilities may induce extremely vigorous mixing through intense three-dimensional turbulence when transition occurs, we focus the discussion here on explaining the mechanisms by which this can occur, concentrating
on recent advances in the explanation and quantification of the time-dependent
mixing behavior obtained through theoretical and numerical analyses. Because
the dynamical processes that occur during the evolution of an unstable free shear
layer feed back in a strongly negative fashion on the height-dependent horizontally

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

137

averaged mean state that supports the initial Ri < 0.25 KH instability, if the initial
mean state is not continuously reinforced the process of shear-layer transition will
have a finite lifetime, the end state of which will be a relaminarized mean state
that is dynamically stable. This is a fundamental point that distinguishes unforced
stratified shear flow evolution from forced stratified turbulence, whether the latter is realized experimentally (see Linden 1979, Fernando 1991 for reviews and,
more recently, Park et al.1994, Keller & Van Atta 2000) or numerically through
the maintenance of the mean shear and density profiles (Gerz et al. 1989, Holt
et al. 1992, Jacobitz et al. 1997, Diamessis & Nomura 2002). As we discuss in
Section 3, several important points of connection do exist between these distinct
circumstances, particularly in flows with sufficiently weak stratification and large
Reynolds number, as in such cases the decay of the turbulence following transition
is slow enough to allow for meaningful comparison.
Analysis of the processes that occur during mixing layer development point
to the utility of consideration of a discrete sequence of interactions, with the
character of the flow passing through an identifiable life cycle. In order to isolate the
fundamental character of these distinct stages, it is most straightforward to assume
that the flow remains horizontally periodic and that the initial instability develops in
time, rather than in space as it does in experiments in which the instability is forced
to develop downstream from a splitter plate (see, e.g., Lawrence et al. 1991). Such
analyses are therefore most directly applicable to understanding the measurements
made in tilted-tube experiments such as those by Thorpe (reviewed in Thorpe
1987). Our specific goal in this review, however, is to discuss the way in which
recent theoretical developments (discussed initially in the papers by Winters et al.
1995 and Winters & DAsaro 1996, which built upon the seminal ideas of Lorenz
1955) are enabling us to make quantitative contact between numerical simulations
of shear flow transition (in particular those by Caulfield & Peltier 2000 and Smyth
& Moum 2000a,b) and experimental observations (Linden 1979, Fernando 1991,
Strang & Fernando 2001). The accumulating data present compelling evidence
that, provided the flow undergoes a true transition to turbulence, the efficiency (i.e.,
the ratio of irreversible increases in the potential energy of the density distribution
to the irreversible loss of kinetic energy) is approximately 20%.
The identification of the sequence of secondary instabilities that control the
evolution of the free shear layer, as well as that for the unstratified end member,
began with the work by Peltier et al. (1978), Davis & Peltier (1979), and Klaassen
& Peltier (1985b, 1989, 1991). In these papers it was first suggested on physical grounds, and then demonstrated mathematically, based on the application of
rigorous Floquet methods, that once the initially two-dimensional stratified KH
instability saturated, it became susceptible to a three-dimensional convective instability that was focused in the regions around the periphery of the vortex cores,
or billows that characterize the cats eye structure of the saturated KH instability
(see Figure 1, where the regions of static instability induced by the primary instability are shaded). The validity of these three-dimensional linear analyses of the
stability of the two-dimensional nonlinear KH billow relied on the assumption of a

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

138

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Figure 1 Contours of spanwise vorticity at a time when the primary KH billow has maximum amplitude in a stratified shear flow with initial hyperbolic tangent velocity and density
profiles. Re = 750, Pr = 1, and minimum gradient Richardson number Ri(0) = 0.05. Statically unstable regions are shaded. [Adapted from Figure 1b in Caulfield & Peltier (2000),
used with permission.]

separation of timescales between the two-dimensional KH billow and the incipient


three-dimensional instability, an assumption that was verified a posteriori.
Three-dimensional direct numerical simulations were required to test the validity of these theoretical predictions, and the analyses described by Caulfield &
Peltier (1994) fully confirmed the prediction that the secondary instability should
consist of a shear aligned convective disturbance focused in regions of overturned
isopycnals caused by the rolling-up of the vorticity in the original shear layer.
The more detailed analyses by Caulfield & Peltier (2000) further demonstrated
the validity of the theoretical prediction that this secondary mode of transition for
stratified flow was fundamentally different from that which controls the transition
to turbulence in the unstratified case. Although the appearance of streamwise oriented streaks of vorticity are observed to be precursory to turbulent collapse in both
cases, in the unstratified case these streaks originate in a hyperbolic instability that
is focused on the hyperbolic stagnation points that exist between successive billow
cores, as first suggested on theoretical grounds by Klaassen & Peltier (1991). This
result contradicted the earlier suggestion that it was the elliptical instability, localized on the elliptical stagnation point interior to the primary KH billow, as had
been earlier suggested by Pierrehumbert & Widnall (1982), that was responsible

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

139

for transition in the unstratified case (see Kerswell 2002 for a thorough review of
the elliptical instability).
As we review in detail here, the efficiency of the mixing that occurs immediately
after the initially two-dimensional nonlinear KH instability matures is rather high,
typically of the order of 70%. Crucially, this mixing is associated with convective
overturning of the statically unstable regions in the periphery of the primary billow,
which trigger the onset of the three-dimensional convective instability, and is
efficient because the flow is in no way turbulent at this stage. Therefore, the
viscous dissipation is relatively close to its laminar value, and thus a significant
proportion of the energy being lost irreversibly from the velocity shear is leading
to irreversible increases in flow potential energy. The relative importance of this
stage is, unsurprisingly, strongly dependent on the flow Reynolds number and
overall stratification, as we show below, and for sufficiently small Richardson
numbers, and sufficiently large Reynolds numbers, the mixing associated with
this stage can be completely overwhelmed by turbulent mixing associated with
the breakdown of the finite amplitude saturated structures owing to the convective
instability.
This point, that the dominant mixing processes for a stratified shear flow are
driven by secondary convective instabilities that are catalyzed by the primary
instability and thus act as the most significant injection scale for the ensuing
turbulent cascade and irreversible mixing, lies at the heart of the discussion to
follow. To make this clear, it is necessary to quantify the relative importance of
mixing events within the dynamically evolving flow as well as the evolution of
flow energetics, in particular, the exchanges between the kinetic energy and potential energy reservoirs. In Section 2, we therefore discuss a framework that may
be employed to achieve this, essentially by compartmentalization of the potential
energy of the flow into two parts, the background or minimum potential energy and
the available potential energy, as suggested by Winters and coworkers (Winters
et al. 1995, Winters & DAsaro 1996), whose methodology builds on the ideas
of Lorenz (1955). Furthermore, this framework is a generalization to three dimensions of a methodology suggested earlier by Thorpe (1977). It is important to
note that a somewhat similar approach was proposed independently by Scinocca
(1995).
Use of this compartmentalization, and appropriately defined mixing efficiencies that may be derived from it, has recently led to a clear understanding as to
why experimental measures of mixing efficiency have invariably led to estimates
on the order of 0.2. In Section 3 we summarize the results of numerical simulations of transitional shear flows, results that point toward the crucial importance
of late-time turbulence transition to understanding the mixing properties of stratified flows at sufficiently high Reynolds number and sufficiently low characteristic
Richardson number. In Section 4 we focus our attention on this late stage of flow
evolution in which transition occurs, demonstrating that mixing efficiency is expected to be a nonmonotonic function of the Richardson number, a result that
has extremely important and well-known dynamical consequences involving the

1 Nov 2002

18:38

140

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

tendency of a turbulent flow of density-stratified fluid to induce a step-discontinuous


staircase structure in the height variation of horizontally averaged density, as
originally proposed by Phillips (1972) and Posmentier (1977) and subsequently
discussed by various authors (Linden 1979, Barenblatt et al. 1993, Balmforth et al.
1998). We discuss the implications of our results in comparison with these recent
theoretical investigations as well as experiments (in particular those by Strang
& Fernando 2001), concentrating on the issues of mixing efficiency and hence
layer formation. In Section 5 we offer conclusions and a discussion of outstanding
questions.

2. BACKGROUND AND AVAILABLE POTENTIAL


ENERGY IN STRATIFIED FLUIDS
The utility of the compartmentalization of the potential energy of a flow into two
parts can best be appreciated by consideration of the evolution equations for the
total flow kinetic energy and potential energy. For simplicity, we concentrate on the
simple circumstance (typically characteristic of numerical simulations) in which
the horizontal boundaries are periodic so that there is no net transport into or out
of the domain. We also assume that the upper and lower boundaries are insulating
so that there is no flux of density into or out of the domain (see Winters et al. 1995
for a discussion of the boundary contributions in more general circumstances).
Because we are interested in stratified shear flows, we assume (as is conventional)
that there is some characteristic total velocity difference 2V0, with characteristic
shear-layer depth 2d0, and total density difference 2 0. We further assume that the
fluid is Boussinesq so that the variations in density are sufficiently small compared
to some characteristic reference density and that they are only significant in the
buoyancy force. Therefore, using 0, d0, and the advective time scale d0/V0, the
nondimensional governing equations (for a fluid with kinematic viscosity and
density diffusivity ) are
Dui
1 2 ui
p0
Rio 0 i3 +
,
=
Dt
xi
Re x2j

(2a)

ui
= 0,
xi

(2b)

1 20
D 0
=
,
Dt
Re Pr x2j

(2c)

where 0 and p0 are the perturbation density and pressure, respectively, Pr is the
Prandtl number /, Re = V0d0/ is the Reynolds number, and Ri0 is the bulk
Richardson number g o d0/V20 , an overall measure of the relative importance of
the stratification compared to the inertia of the flow. With the appropriate boundary
conditions described above, the evolution of the kinetic energy K per unit volume

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

STRATIFIED SHEAR FLOW MIXING

P1: IBC

141

of the flow governed by Equations 2ac satisfies the equation


1
dK
= Rio hhhwix iy iz
hhh(u)2 ix iy iz
dt
Re

(3a)

= H ,

(3b)

K = hhh(u2 + v2 + w2 )/2ix iy iz ,

(3c)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

with

where angle brackets denote averaging over the subscripted coordinate direction.
Similarly the potential energy P of the flow governed by Equations 2ac satisfies
the evolution equation
dP
= H + Dp ,
dt

(4a)

P = Rio hhhzix iy iz

(4b)

with

and
Dp =

Rio ( hbottom, ti (top, t))


,
Z Re Pr

(4c)

where Z is the nondimensional height of the domain, and (bottom) and (top)
are the nondimensional densities at the bottom and top boundaries, respectively.
Dp is the rate at which the potential energy of a statically stable density distribution would increase in the absence of macroscopic fluid motion (essentially
through conversion of internal energy into potential energy), and so for sufficiently
deep flow domains this plays no significant dynamical role in flow evolution. We
wish to distinguish this increase from mixing because we prefer to consider mixing as involving irreversible changes in fluid properties that are directly associated
with macroscopic fluid motions, rather than those changes that would occur in a
diffusive, macroscopically quiescent, system.
The crucial term for the quantification of mixing in these equations is, of course,
the heat flux, or buoyancy flux H, which quantifies the exchange between the kinetic and potential energy reservoirs. However, fluid motions can increase the
potential energy of the system by two qualitatively different processes, namely,
by either mixing or stirring. Any lifting (on average) of anomalously dense
fluid by the flow (hence involving negative values of H) will lead to an increase
of potential energy. However, this increase is only irreversible (and so is appropriately identified with mixing) if diffusion acts on sufficiently small scales so as to
change the density of the lifted fluid parcel, as well as its neighbors. Such mixing
is inevitably irreversible and acts over the small length scales characteristic of diffusion. Otherwise, the lifted parcels may later fall under gravity, reconverting the

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

142

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

temporarily stored potential energy into kinetic energy, a process that is identified
as stirring (a larger-scale process that is nondiffusive).
It should be clear, therefore, that instantaneous values of the buoyancy flux H
can be very misleading as a proxy for mixing, particularly in flows with significant
time variability. Irreversible increases in potential energy owing to fluid motion can
only be distinguished from the buoyancy flux through averaging over sufficiently
long timescales so that the reversible effects of stirring completely cancel out.
Clearly, such averaging is not always possible in the laboratory or in the analysis
of oceanographic or meteorological data. However, the formalism of Winters et al.
(1995) does enable the time-dependent separation of mixing from stirring processes
because variations in the background potential energy can be identified directly
with mixing and the diffusion of the mean profile quantified by Dp (which these
authors have referred to as 8i).
We define the dimensional background potential energy P B(t) as
PB (t) = hgB (z, t) ziz .

(5)

In Equation 5, B(z, t) is the notional one-dimensional density distribution generated (by volume conserving adiabatic rearrangement of fluid elements) from the
actual density distribution (x, y, z, t) at the time t that is everywhere statically
stable. In general, the background potential energy is the minimum potential energy
that the flow can have, and this energy cannot be extracted from the potential
energy reservoir to drive (macroscopic) fluid motions, i.e., energy stored in P B
is inaccessible to stirring. Following Lorenz (1955), we then define an available
potential energy P A that is, at least in principle, available to the kinetic energy
reservoir for stirring as
PA = P PB ,

(6)

where P is the instantaneous total potential energy of the flow as defined in Equation 4b.
P A > 0 if there exists any lateral variations in density or statically unstable
regions. In Figure 2, we provide a schematic illustration of just such an adiabatic
rearrangement of fluid elements, showing that the generation of B can involve
changes in shape, but not the volume of the individual elements, and that the total
potential energy (defined by the mean density profile) is always bounded below
by the background potential energy.
In the absence of macroscopic fluid motion, P B will increase through the effect
of Dp, as defined in Equation 4c, as the mean profile diffuses. Crucially, P B can also
increase through irreversible mixing processes, and thus it is possible to partition
the buoyancy flux into two distinct terms and to write distinct evolution equations
for P A and P B as
dPA
= H M,
dt

(7a)

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

143

Figure 2 Schematic diagram of the adiabatic redistribution of fluid parcels (with densities
in the range +/ 31) to evaluate the background density distribution B associated with
the minimal background potential energy P B of a flow. Note that the fluid parcels change
shape but not volume. Also shown are the vertical distributions of the mean density and
the background density distribution B. [Adapted from figure 3 in Caulfield & Peltier (2000),
used with permission.]

dPB
= M + DP ,
dt

(7b)

thereby defining an irreversible mixing rate M. The stirring rate is then S =


H M. Over sufficiently long time integrations, the integral of the stirring rate
is zero. Winters et al. (1995) referred to the sum of the terms on the right-hand side
of Equation 7b as 8B. In our view, it is most appropriate to consider the mixing
rate M and the term Dp to be distinct, as this subdivision isolates the component
of the increase in P B that originally came from the kinetic energy reservoir and
thus is inherently related to macroscopic fluid motions. When there is fluid motion,
kinetic energy enters the potential energy reservoir through the buoyancy flux H,
leading to increases (at least briefly) in the available potential energy P A. Mixing
(increases in P B) is associated with an imperfect return of this potential energy
to the kinetic energy reservoir, so for mixing to occur it is necessary that there be
some available potential energy in the system.
For fluid motions to enhance diffusive transport, and hence the rate of increase
of P B, two distinct phenomena must be involved. First, local gradients of density

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

144

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

must be intensified to enhance the flux. However, this is only half the story: In
order to enhance transport, the area over which the flux enhancement occurs must
also increase, implying the need for neighboring density isosurfaces to become
complicated through the development of small scales of motion. A dense parcel,
feeling a buoyancy force restoring it to its neutral buoyancy height, will tend to
cause both these effects: Local gradients in density will intensify owing to the parcels motion (thus enhancing diffusive flux), and also, critically, the conversion of
potential to kinetic energy will tend to introduce small scales, thus enhancing transport. As we discuss below, such behavior can be directly identified and understood
in transitional shear flows but can also be found in forced stratified turbulence, provided that the flow exhibits regions of overturning (e.g., see Diamessis & Nomura
2002).
Indeed, a particular attraction of this formalism is that it is exceedingly straightforward to implement in a numerical simulation, as suggested by Winters et al.
(1995). At any particular time step, sorting algorithms can be used to order by
their density the n individual fluid elements located in the n grid volume elements
in the computational domain. These fluid parcels can then be notionally placed
into a purely one-dimensional array, created by deforming the n grid volume elements (while conserving volume) into n thin horizontal layers extending across
the whole flow domain. [See Caulfield & Peltier (2000) for a more detailed discussion, generalizing Winters et al. (1995), who proposed using the nondeformed
grid, thus allowing for horizontal density variations and hence generating only an
approximate upper bound to B.] This notional deformed grid, corresponding to
the identification of a unique height with each of the fluid elements arrayed within
a monotonic density field, creates the background density profile B(z) and hence
P B. As originally pointed out by Winters et al. (1995), and discussed further by
Tseng & Ferziger (2001), such sorting and reordering is intimately related to construction of the probability density function for the fluid density. Numerically, this
procedure can be performed at every time step, and so the time evolution of P B
can be followed, thus identifying the time dependence of the irreversible mixing
that is occurring.
It is of course important to quantify not only the absolute amount of mixing that
occurs, but also its efficiency, i.e., the proportion of the kinetic energy lost by the
fluid that leads to mixing, as opposed to that which is lost to viscous dissipation.
In the discussion to follow, we make use of a definition of instantaneous mixing
efficiency, say, E i, that we define as

Ei =

M
.
M+

(8a)

Because the viscous dissipation is positive-definite, it is clear that E i 1, as


is required on physical grounds. The denominator here represents the irreversible
losses of kinetic energy owing to both irreversible mixing and viscous dissipation.

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

STRATIFIED SHEAR FLOW MIXING

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

We may also define a cumulative mixing efficiency in the obvious way as


Rt
0 M(u) du
,
Ec = R t
Rt
0 M(u) du + 0 (u) du

145

(8b)

where the time interval may be selected so as to isolate the manner in which
mixing efficiency changes over a series of appropriately selected epochs during
flow evolution. Such a definition eliminates the diffusive contribution DP, as this
is not associated with the kinetic energy, and hence the macroscopic fluid motion
of the system.
This elimination allows an identification of the cumulative mixing efficiency,
integrated over a sufficiently long time, with the flux Richardson number as was
defined by Linden (1979), namely, as the fraction of the change of the available
kinetic energy which appears as the potential energy of the stratification. This
is appropriate because it allows for the long-term cancellation of the influence of
those large-scale processes that we refer to as stirring. However, the flux Richardson
number Rf is often defined as
Rf =

H
.
H +

(9a)

or, indeed, in sheared flows as


Rf =

H
u0 w0 dV/dz

(9b)

(see, e.g., Strang & Fernando 2001), these two definitions being equivalent if
the system is closed, i.e., the turbulent kinetic energy is statistically steady and
dominates the dissipation occurring within the flow.
The flux Richardson number is a parameter that lies at the heart of the parameterizations of mixing within turbulent closure schemes (see, e.g., Mellor & Yamada
1982). Here, we are interested in the mechanisms responsible for causing, and the
quantification of, stratified mixing within transitional stratified shear flows, and
in such fundamentally time-dependent circumstances, Rf would be an adequate
definition of mixing efficiency if the entire increase of potential energy caused by
the vertical flux of buoyancy H were irreversibly converted into an increase of
background potential energy P B. However, it is inevitable that some (often very
significant) fraction of H is simply associated with reversible stirring, and so it
should be clear that our definitions of mixing efficiency are most appropriate when
the required data are available, as is the case in numerical simulations.
Our procedure for the determination of mixing efficiency in stably stratified
parallel flows reviewed above is essentially a three-dimensional generalization of
the one-dimensional reordering proposed by Thorpe (1977) to assess the size of
overturning regions from one-dimensional observational profiles, generating the
so-called Thorpe scale LT. His scale is based on the reordering of individual density
measurements in such a way as to generate a stable vertical profile. Such sorting
enables one to generate an array of Thorpe displacements j for each of the

30 Nov 2002

9:43

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

146

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

density parcels. From these, one defines the Thorpe scale LT = ( 2j )1/2 as the rms
of the displacements. Flows in which mixing is intense are then flows in which
the Thorpe scale is large. Smyth & Moum (2000b) have examined in detail the
connection between the Thorpe scale and the characteristic scales of reordering
within a three-dimensional framework.
As described below, construction of the subdivision of the potential energy into
P B and P A, together with a consideration of the time-dependent evolution of these
fields, the quantities M and S = H M, and the mixing efficiencies defined
in Equation 8, yield valuable insight into the nature of the mixing that occurs in
stratified shear flows.

3. THE ANATOMY OF THE MIXING TRANSITION IN


STRATIFIED PARALLEL FLOW
In reviewing recent work on the detailed sequence of dynamical events that occurs during the process of transition from a laminar state of stratified parallel
flow to a state of intense three-dimensional turbulence, it is useful to discuss the
various regimes of flow sequentially. There are essentially three distinct phases
in the evolution of such flows: (a) the stage of initial parallel shear instability and subsequent KH wave growth to saturation amplitude, (b) the stage of
onset and growth of the three-dimensional secondary instability through which
shear aligned convective rolls are nucleated in the region surrounding the KH
billow cores, and finally, (c) the stage of turbulent-mixing layer collapse characterized by intense viscous dissipation and triggered fundamentally by the breakdown of the shear aligned convective rolls that grow on, and then destroy, the
primary KH billow. These three stages of the mixing transition are illustrated
qualitatively in Figure 3, where we show isosurfaces of cross-stream and streamwise vorticity from a direct numerical simulation (DNS) in which the streamwise length of the domain is restricted to a single wavelength of the fastest
growing mode of the linear instability through which the transition process is
initially engendered. Below, we discuss these stages in the mixing process in
sequence.

3.1. The Kelvin-Helmholtz Stage of Flow Evolution


To illustrate through specific examples the sequence of events that occurs during
the transition to turbulence in the stratified shear layer, we consider an initial state

characterized by background profiles of horizontal streamwise velocity V(z)


and
density (z)
in the form

V(z)
= Vo tanh (z/d),

(10a)

(z)
= a + o tanh (Rz/d),

(10b)

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

147

in which R = 1.1 is the ratio of the thickness of the shear layer to the thickness of the
density inversion and is chosen to enable close accord with conditions achievable in
laboratory experiments (e.g., Thorpe 1987) and to be susceptible to instabilities of
KH type. This flow has been considered by many authors (Caulfield & Peltier 1994,
2000; Palmer et al. 1994, 1996; Cortesi et al. 1998; Smyth & Moum 2000a,b). We
briefly review what we perceive to be the key characteristics of the transition to a
stage of intense turbulent mixing in these flows, focusing on the critical importance
of the secondary convective instability to flow evolution as well as the variation of
the mixing characteristics with stratification. For these profiles, the initial gradient Richardson number is minimum at the midpoint of the shear layer and equals
Ri(0) = R Ri0, where Ri0 is the bulk Richardson number of the initially parallel
flow. We discuss eight different simulations in detail here, all with Pr = 1; Re =
750; and Ri(0) = 0, 0.025, 0.05, 0.0625, 0.075, 0.0875, 0.1, and 0.125. This
Reynolds number is comparable with that usually realized in laboratory experiments. Such a detailed investigation of the isolated effect of the variation in stratification enables us to identify clearly the nature of the nonmonotonicity of the
variation of the mixing efficiency with stratification, although it is important to
stress that it is also of great interest to understand the effect of varying the other
characteristic flow parameters (see Klaassen & Peltier 1985a; Smyth & Moum
2000a,b; Smyth et al. 2001 for a discussion of the effect of variation in Pr). We
return to this point in Section 5.
We have also chosen the density stratification to vanish in the region exterior
to the region of strong shear. When the exterior region is also stably stratified,
internal waves may be emitted from the region of turbulent collapse and the nature
of the evolution of the turbulent flow may be significantly affected. This interesting
source of complexity is not addressed here (see Sutherland & Peltier 1994, 1995;
Sutherland et al. 1994; Werne & Fritts 1999 for consideration of the behavior of
flows in which the influence of external stratification is included).
For all the flows that we consider, the primary KH instability develops in
a completely two-dimensional manner, with an initial exponential growth rate
that is well predicted by linear theory and that eventually saturates at nondimensional time t 4060, with this saturation time being dependent on initial
stratification. As discussed in detail by Caulfield & Peltier (2000), the roll-up
of the primary billow leads to significant increases in the available potential
energy P A, which is associated with the development of the statically unstable regions at the periphery of the primary billow (as is apparent in Figure 1).
There also is typically little irreversible mixing before the primary billow saturates. At this saturation time, the primary billows are usually very coherent,
although the outer-scale Reynolds number (defined by the vertical extent of the
billow and the total velocity jump) is on the order of 104 and so is sufficiently
high for us to expect a turbulence transition (see Dimotakis 2000 for further discussion). This is naturally due to the evolution of the flow in two dimensions:
As the flow becomes three dimensional, its character changes qualitatively (see
below).

1 Nov 2002

18:38

148

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

3.2. The Stage in Which Secondary Instability of the


Kelvin-Helmholtz Structure Occurs
Indeed, as the two-dimensional KH billows saturate, they become susceptible to
disruption by a family of three-dimensional secondary instabilities that begin to
grow precisely at the instant when the KH billow reaches its maximum amplitude.
The stratified KH billow is not a stable flow configuration, owing critically to the
significant value of the available potential energy P A, which is associated with
the convectively unstable regions at the periphery of the primary billow cores.
The properties of these secondary instabilities may be accurately deduced on the
basis of a detailed Floquet analysis. Subject to the assumption that the threedimensional perturbations to the two-dimensional nonlinear KH billow remain
small, we represent them in the form
u 3D (x, y, z, t) = u 3D (y, z, t) exp [i x + iy],

(11a)

3D (x, y, z, t) = 3D (y, z, t) exp [i x + iy],

(11b)

where x is the spanwise direction and y is the streamwise direction. Restricting


attention to fluctuations of streamwise (y) wavenumber equal to or greater than
the wavenumber KH of the two-dimensional KH instability, we set the Floquet
exponent to zero. The spectrum of subharmonic merging instabilities has been
discussed in detail by Klaassen & Peltier (1989). As shown in the simulations by
Cortesi et al. (1998) and Smyth & Moum (2000b), the transition to turbulence
within a stratified shear flow is an inherently three-dimensional process, with twodimensional merging events still yielding a highly coherent state, that is destroyed
by the development of three-dimensional structures that such detailed stability
analyses can identify.
Subject to the assumption that the three-dimensional secondary instabilities
grow so rapidly that the two-dimensional KH billow may be assumed to be time
independent, the dependence of u 3D and 3D on space and time becomes separable.
We may therefore assume
u 3D (y, z, t) = u 3D (y, z)e t ,

(12a)

3D (y, z, t) = 3D
(y, z)e t ,

(12b)

with, in general, the complex growth rate of


differential equations satisfied by the fields u3D
perturbation pressure p3D , in which u2D = (0,
locity and density fields of the two-dimensional
follows:
u 3D + v2D

= r + ii . The linear partial

= (u3D , v3D , w3D ), 3D


and the
v2D, w2D) and 2D are the veKelvin-Helmholtz wave, are as

1
u 3D
u

+
+ w2D 3D = i p3D
D2 u 3D
y
z
Re

(13a)

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

149

STRATIFIED SHEAR FLOW MIXING

v3D + v2D

p
1
v3D
v
v2D
v2D
+ w2D 3D + v3D
+ w3D
= 3D +
D2 v3D
y
z
y
z
y
Re
(13b)

w3D + v2D

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

+ v2D
3D

w3D
w
w2D
w2D
+ w2D 3D + v3D
+ w3D
=
y
z
y
z

1
p3D

+
Rio 3D
D2 w3D
z
Re

(13c)

1
3D

2D
2D

+ w2D 3D + v3D
+ w3D
=
D2 3D
y
z
y
z
Re Pr

i u3D +

v3D
w3D
+
= 0,
y
z

(13d)
(13e)

where the operator D2 = 2 + 2 / y 2 + 2 /z 2 . Because the pressure perturbation may be determined on the basis of the following diagnostic equation,
"
D2 p3D

= 2

v2D
y

w2D
+
y

v3D
y

v3D
z

w2D
z

#
Rio

w3D
z

v2D
z

w3D
y

3D
,
y

(13f)

we may eliminate p3D from the set (Equation 13). Thus, u3D is also removed,
thereby reducing the eigenvalue problem for to one involving only v3D , w3D , and

3D
. This system may be solved using a Galerkin method, which has the advantage
of rapid convergence.
To implement the Galerkin methodology, we expand the unknown fields on a
basis of orthogonal functions chosen such that the boundary conditions are satisfied
automatically. Specifically, we take


=
v3D , w3D , 3D

+ X
X

(Amn cos[n (z H)/H], Bmn sin[ ], Cmn cos[ ])eimKH y .

(14)

m= n=o

Substitution of these expansions into Equations 13ae, followed by computation


of the inner product of each of the equations in the set by each of the members of
the Galerkin basis in turn, delivers the following set of algebraic equations for the
Galerkin amplitudes Amn, Bmn, and Cmn:

30 Nov 2002

9:44

150

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD
mn
mn
Apq = hAAimn
pq Amn + hABipq Bmn + hACipq Cmn ,

(15a)

mn
mn
Bpq = hBAimn
pq Amn + hBBipq Bmn + hBCipq Cmn ,

(15b)

mn
mn
Cpq = hCAimn
pq Amn + hCBipq Bmn + hCCipq Cmn .

(15c)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

and

Expressions for the four-dimensional coefficient arrays hAAi, hABi, etc., are provided by Smyth & Peltier (1994). By imposing a particular truncation scheme
upon the Galerkin expansions (Equation 14), specifically we use 2|m|+ n N,
and by taking N = 41 so that adequate resolution is achieved, Equation 15 then
poses a standard matrix eigenvalue problem for the eigenvalue . A single eigenvector is defined simply by concatenating the unknowns Apq, Bpq, and Cpq into a
single vector V with a single running subscript. Potylitsin & Peltier (1998, 1999)
have illustrated the eigenfunctions obtained in such analyses, including the impact
of background rotation parallel to the axis of the KH vortex tubes, by directly
contouring the three-dimensional perturbation vorticity field.
In Figure 4, we plot the real part of the eigenvalue , the growth rate r, as a
function of the spanwise wavenumber of the fastest growing three-dimensional
perturbation for Ri(0) = 0, 0.025, 0.05, and 0.1. On each diagram, the globally
most unstable mode is shown as the open circle. In the unstratified case, the growth
rate for the subdominant elliptical core mode is also indicated. Figure 5 shows the
cross-stream average over a wavelength 2/ of the perturbation kinetic energy
K3D determined on the basis of the eigenfunction of the most unstable mode from
K3D =

1 2
|u 3D | + |v3D |2 + |w3D |2 x .
2

(16)

Inspection of these results demonstrates that the fastest growing mode of threedimensional secondary instability becomes more sharply confined to the statically
unstable periphery of the individual billow cores as the strength of the ambient
stratification increases. Only in the limit where the background stratification vanishes entirely does the mode become focused on the hyperbolic stagnation points
that are located between adjacent billow cores. In Figure 5a, we show not only
K3D for the hyperbolic mode, but also K3D for the elliptical core mode by Pierrehumbert & Widnall (1982), the latter denoted by the heavy solid contours in the
billow core of the unstratified flow.
That these predictions of the linear stability analysis of finite amplitude KH
billows are precisely borne out by complete DNS solutions of the initial value
problem is demonstrated in Figure 6, where isosurfaces of spanwise vorticity
and streamwise vorticity from two typical three-dimensional DNS integrations
over the range of time during which the streamwise vortex streaks have just
emerged from the background two-dimensional KH flow of finite amplitude are
shown. Results for both the unstratified flow and the stratified flow for which

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

151

Figure 4 Predicted growth rate r of the most unstable three-dimensional mode against
spanwise wavenumber for Ri(0) = (a) 0, (b) 0.025, (c) 0.05, and (d ) 0.1. In each case, the
globally most unstable mode (and for the unstratified flow, the most unstable core-centered
mode) is marked with a circle.

the strength of the stratification is characterized by Ri(0) = 0.05 are also shown.
Intercomparison of the results for the unstratified flow (top) and the stratified flow
(bottom) demonstrates that the distinctly different regions of localization of the
secondary instability that is predicted by the Floquet analysis is precisely borne
out by the DNS results. In the unstratified case, streamwise vortex streaks originate through instability focused on the hyperbolic stagnation points that separate
adjacent billow cores, whereas in stratified flow, these streaks originate through
convective instability in the regions of overturned isopycnals. Their development
is triggered by a single convective overturning (releasing available potential energy
to the kinetic energy reservoir) in the statically unstable regions at the periphery of
the primary cores. Their growth is also driven by the background shear, with the
relative importance of these two processes depending on the overall stratification
(see Caulfield & Peltier 2000 for further discussion). Furthermore, the critical importance of these streamwise vortices for the transition to turbulence is confirmed

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

152

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Figure 5 Contours of K3D (as defined in Equation 16) for the most unstable threedimensional mode in flows with Ri(0) = (a) 0 (thin solid lines show the most unstable
(hyperbolic) mode with = 2.5, and thick solid lines show the most unstable elliptical mode
with = 0.9); (b) 0.025; (c) 0.05; and (d ) 0.1. [Adapted from figure 7 in Caulfield & Peltier
(2000), used with permission.]

by these simulations, where the primary billows were so strongly disrupted by


these secondary instabilities that transition occurred before subharmonic merging
could take place.
The time-dependent evolution of cross-stream power spectra of the flow may
be usefully invoked to test the quality of the further linear theoretic prediction
of the expected cross-stream wavenumber of the three-dimensional secondary
instability that is responsible, ultimately, for inducing complete turbulent collapse
of the shear layer. Because the boundary conditions in the x-direction employed
for the purpose of the DNS analyses are periodic, we may represent the streamwise
component of the vorticity in the form of a Fourier series as

x (x, y, z, t) =

+N
X
n=N

Cn (n , y, z, t)ei nx ,

(17a)

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

153

Figure 7 Variation of the various components of the normalized power spectrum density
P( n, t) (defined in Equation 18) against n at four characteristic times within a flow with
Ri(0) = 0.05. The wavenumber of the most unstable spanwise mode shown in Figure 4 is
denoted by a thick vertical dashed line.

with

Z
Cn (n , y, z, t) =

Lx

x ei nx dx

(17b)

and n = n(2/Lx ), where Lx is the cross-stream width of the domain. Figure 7


shows the domain integral of the normalized spanwise power spectrum
P(n , t) = hh|c(n , y, z, t)|2 iy iz .

(18)

Inspection of the results in Figure 7, on which the prediction of Floquet theory


of the cross-stream wavenumber of the fastest growing mode of linear theory is
shown as the vertical dashed line, demonstrates that this prediction of linear theory
is precisely recovered by the detailed DNS analysis, thus verifying the assumption
of a separation of timescales on the basis of which the stability analysis of the
nonlinear KH waves was performed. By the time of the final frame of Figure 7

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

154

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

(t = 156 in nondimensional units), the power spectrum of the streamwise vorticity


has assumed the red form that one expects of a strongly turbulent flow in which
a cascade to small scales occurs following injection at the scale of the crossstream mode of secondary instability. It is critical to appreciate that this scale is
substantially smaller than the primary billow scale (by approximately one order
of magnitude), and this has important implications for the mixing within the flow
(discussed below). Indeed, it is clearly apparent that transition has occurred, as is
expected because the flow Reynolds number is sufficiently high (see Dimotakis
2000). This is strong evidence to suggest that the crucial physical structures that
trigger turbulence are not the primary billows, but rather the secondary convective
instabilities that develop into streamwise vortices. The primary billows should be
thought of as acting as catalysts, creating a flow that is conducive to the development of convective instabilities that develop to finite amplitude and then trigger
transition. Important questions to address concern the mechanism or mechanisms
by which the inherently three-dimensional streamwise vortices trigger transition as
well as the mixing characteristics associated with the flow evolution, in particular,
subsequent to this transition.

3.3. The Dominant Stage of Intense Three-Dimensional


Mixing of the Stratified Parallel Flow
Kinetic energy diagnostics are shown in Figure 8 for four representative threedimensional DNS. Figures 8a,b, and c, respectively, present the evolution of total
kinetic energy of the flow K, the cross-stream average perturbation kinetic energy
of the flow KKH, defined such that
hKix = K + KKH ,

(19a)

K = h(V 2 )/2iz ,

(19b)

with

and
KKH =

v2KH + w2KH


2 y z,

(19c)

and the three-dimensional deviation from this cross-stream average, which we are
denoting by K3D and which is defined as
K3D = K K KKH .

(19d)

This compartmentalization allows us to isolate the evolution of three dimensionality from the evolving two-dimensional KH billow component of the flow.
Inspection of these diagnostics will demonstrate (from Figure 8a) that it is not
until nondimensional time t > 100 that total kinetic energy begins to decrease
rapidly owing to the intense viscous dissipation that onsets only with the proliferation of small spatial scales that develop during three-dimensional turbulent

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

155

Figure 8 Time evolution of (a) K(t)/K(0), (b) KKH(t)/K(0), and (c) K3D(t)/K(0) for threedimensional (as defined in Equations 3 and 19) simulations with Ri(0) = 0 (dotted line),
Ri(0) = 0.025 (solid line), Ri(0) = 0.05 (dashed line), and Ri(0) = 0.1 (dot-dashed line). The
characteristic times given in the table are marked with a circle [Ri(0) = 0], a plus [Ri(0) =
0.025], a cross [Ri(0) = 0.05], or an asterisk [Ri(0) = 0.1]. [Adapted from figure 8 in
Caulfield & Peltier (2000), used with permission.]

transition and shear-layer collapse. Furthermore, the onset of three-dimensional


perturbation growth is exactly coincident with the saturation of the primary KH
billow, reinforcing the important conceptual idea that the two-dimensional billow
serves merely to catalyze the development of three-dimensional motions. On each
of these kinetic energy diagnostic time series, we have denoted by appropriate
symbols three significant times in the evolution of the flow, respectively: the time
t2Dmax at which the two-dimensional KH instability achieves maximum amplitude;
the time td at which the dissipation starts to grow markedly, signaling the onset of
transition; and the time t3Dmax at which the three-dimensional perturbations achieve
maximum amplitude.
Figure 9 (ag) illustrates the evolution of each of the reservoirs of potential
energy (P, P A, and P B) that accompanies the previously discussed evolution of
the various components of kinetic energy. Results are shown for each of the seven

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

156

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Figure 9 Time variation of P (t)-DPt-P (0) (solid line) (as defined in Equation 4), P B(t)-DPtP B(0) (dashed line) (as defined in Equation 5), and P A(t) (dotted line) (as defined in Equation
6) for inherently three-dimensional simulations with Ri(0) = (a) 0.025, (b) 0.05, (c) 0.0625,
(d ) 0.075, (e) 0.0875, ( f ) 0.1, and (g) 0.125, compared with the time variation of E i (solid
line) and E c (dashed line) (as defined in Equations 8ab) for inherently three-dimensional
simulations with Ri(0) = (h) 0.025, (i) 0.05, ( j) 0.0625, (k) 0.075, (l ) 0.0875, (m) 0.1, and
(n) 0.125. The characteristic times shown in Figure 5 are marked with thick vertical dotted
lines.

stratified flows. The three key times discussed above in connection with the kinetic
energy diagnostics are shown on each frame as the vertical dotted lines.
Directly after the saturation of the primary billows, for each simulation there
is a phase of early time, preturbulent [using the nomenclature by Smyth et al.
(2001)] mixing that is extremely efficient. The instantaneous efficiencies, which
maximize subsequent to the time the two-dimensional waves saturate, reveal the
influence of the nutation and oscillation of the billow core that follows saturation.
Instantaneous mixing efficiency is observed to equal or exceed the high value of
0.75 near the time that the KH wave achieves maximum amplitude even though
the flows remain highly coherent. It is apparent that P B only starts to increase
once P A is large, and indeed much of the increase in P B occurs when the total

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

157

potential energy is decreasing (and so H > 0 as defined in Equation 3). Thus the
nave definition of the flux Richardson number given by Equation 9a is not useful.
At these Reynolds numbers, this early-stage mixing is associated with the convectively unstable regions around the periphery of the primary vortex core, which
naturally have the required characteristics for enhanced mixing (as discussed in
Section 2), i.e., regions of static instability associated with a source of P A, and
locally intensified density gradients. At sufficiently high Reynolds number, this region of the flow experiences enhanced small-scale mixing, as the available potential
energy is imperfectly converted back into kinetic energy. This mixing is primarily
due to the convective overturning that triggers the growth of the three-dimensional
perturbations, in the regions around the periphery of the primary billow core (see
Caulfield & Peltier 2000, where this identification could be made by the analysis
of spatially compartmentalized three-dimensional perturbation growth), and nucleates the characteristic streamwise vortices, which, at sufficiently high Reynolds
number, qualitatively change the long-time behavior of the flow.
Indeed, particularly for flows with weaker initial stratification, the total potential energy increases significantly at the same time as the total kinetic energy
begins to decrease significantly (compare with Figure 8). It is important to note
that this increase in potential energy at late time is not primarily due to a reversible
exchange between the kinetic energy and the available potential energy reservoirs.
Rather, the available potential energy P A exhibits strong variability (necessary to
enable irreversible mixing to occur) but of relatively small amplitude. The observed strong monotonic increase of total potential energy, at times later than the
second characteristic time at which the dissipation starts to increase markedly, is
clearly due to rapid increase in the background potential energy P B of the flow,
a further indication that it is in this latter stage of flow development in which
significant three-dimensional irreversible mixing is occurring. In fact, enhanced
three-dimensional irreversible mixing continues unabated during the entire posttransitional stage of flow development.
An important question to address clearly concerns the reason why this turbulence transition occurs. Once the convective instability develops, a large body of
numerical evidence (see Palmer et al. 1996; Cortesi et al. 1998; Werne & Fritts
1999; Caulfield & Peltier 2000; Smyth & Moum 2000a,b; Smyth et al. 2001) points
toward the inevitable development of disordered turbulent flow, with the streamwise vortices being driven to finite amplitude principally through vortex stretching
by the mean shear, with the primary billows once again playing a primarily catalytic role.
It is apparent that the transition to turbulence is triggered by the catastrophic
breakdown of these streamwise vortices. Owing to their location, the advection
and amplification through stretching of the streamwise vortices by the background
shear leads inevitably to the streamwise vortices propagating toward the braid
region, and there comes an instant when these vortices collide. Such a finite amplitude subcritical interaction occurred in all the simulations analyzed here and
appears to be the inevitable precursor to the onset of truly turbulent motions,

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

158

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

reinforcing the concept that it is the characteristic scale of the streamwise vortices,
and not the significantly larger primary billow scale, that is the dominant injection
scale for the turbulent cascade. Indeed, owing to the unsteadiness of the flow, it
appears not to be appropriate to consider the development of tertiary instabilities
of the streamwise vortices to be the trigger for transition. Rather, subcritical, finite
amplitude vortex-vortex interactions seem to be the primary cause. Indeed, the
numerical evidence appears to point to the requirement not only of adequate resolution of individual streamwise vortices, but also of sufficient spanwise expanse
so that multiple streamwise vortices can develop and interact during transition in
order to capture properly the transition behavior.
However, it is important to stress that there is some evidence (Arendt et al. 1997,
Fritts et al. 1998) that the streamwise vortices may support a family of propagating
disturbances, referred to as Kelvin twist waves (see Arendt et al. 1997, Fritts et al.
1998 for a full discussion). These disturbances, which propagate along the vortex
tubes [and appear to arise naturally as a part of any initial disturbance on the
vortex (as demonstrated by Arendt et al. 1997)], can lead to fragmentation and
depletion of the vortex tubes, thus triggering transition to smaller and smaller
scales of motion within the flow. Arendt et al. (1997) and Fritts et al. (1998) also
found that vortex-vortex interactions with longer wavelengths [associated with the
cooperative Crow (1970) instability of neighboring elliptical vortices] could also
induce turbulence, and it is not yet clear which of these competing mechanisms,
if either, plays an active role in the transition process.
However, it is clear that the turbulence transition is intimately related to the
streamwise vortices induced by convective instability, as first described by Klaassen
& Peltier (1985b), and that they are the primary mechanism by which enhanced
dissipation is triggered within stratified shear flows. The issue as to whether this
turbulence transition mechanism is the dominant mechanism in all stratified shear
flows is highly complex, with the relative importance of the various stages of flow
evolution depending on both the flow parameters and the strength of the initial
forcing. For example, if the flow is forced so that merging events can occur before
the transition to turbulence, the mixing associated with the coherent, and essentially two-dimensional, merging of neighboring vortices can constitute a dominant
part of the total mixing experienced within the flow, as discussed by Smyth et al.
(2001).
Also, and as is apparent for the more strongly stratified simulation [with
Ri(0) = 0.125], the mixing associated with the preturbulent overturning in the
convective layers is very intense and constitutes a significant proportion of the
total mixing experienced by this flow. As is apparent in Figure 8b, the growth of
the primary KH billow in this flow is relatively weak, owing to the stabilizing
effect of stratification on this simulation. Also, as is apparent in Figure 8c, increasing stratification also reduces the growth rate of the developing three-dimensional
perturbations, and their saturation amplitude also reduces as the stratification increases. Such weakening growth of the primary KH instability for Ri(0) > 0.1
is commonly observed in laboratory experiments (see Koop & Browand 1979,

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

159

Linden 1979, Thorpe 1987, Fernando 1991). However, the crucial influence of
sufficiently high stratification actually appears to be the reduction in the saturated
amplitude of the three-dimensional perturbations, which, we believe, is indicative
of a qualitative change in the character of the flow.
Such behavior is also observed in numerical simulations of flows with constant
shear and density gradient, which are likely to have certain points of connection
with the turbulent flows considered here because the characteristic scales of motion
are becoming small compared to the mean shear and density gradient. As identified
by several authors (Holt et al. 1992, Jacobitz et al. 1997, Diamessis & Nomura
2002), the growth of turbulent kinetic energy is inhibited by buoyancy forces for
Ri > 0.1 in such flows. Our results are strongly suggestive of an analogous behavior,
with a relatively short period of turbulent disordered motion occurring for the flows
with higher Ri(0), whereas the less strongly stratified simulations exhibit extended
periods of turbulent mixing, leading to much greater total mixing than occurs in
the initial preturbulent phase. The relative importance of these two phases is likely
to also depend on the Reynolds number of the flow, as lower Reynolds number
will tend to reduce the significance of the turbulent phase at a given stratification
from the high Re behavior that shows a demarcation around Ri(0) = 0.1.
Nevertheless, in all the simulations we discuss here, it is clear that after transition occurs the absolute value of the background potential energy begins to grow.
This is particularly noticeable for flows with smaller initial stratification, where the
preturbulent convective overturning has achieved little mixing, yet the streamwise
vortices that are induced by this overturning grow rapidly and then trigger significantly increased mixing and strong dissipation. This behavior, as is apparent from
the spectra shown in Figure 7, occurs at substantially smaller scale than both the
primary KH instability and the injection scale of the secondary convective rolls,
thereby further reinforcing the analogy with the constant shear/constant density
gradient simulations.
Based on the use of Equations 8a and 8b, we show in Figure 9hn the evolution of the mixing efficiency, both instantaneous (solid lines) and cumulative
(dashed lines), for each of the seven stratified flow examples for which we have
performed detailed analysis. In this range of time during which three-dimensional
motions have proliferated and intense viscous dissipation is occurring, a state that
is turbulent in the conventional sense, the instantaneous mixing efficiency has
decreased substantially toward the experimentally determined value of approximately 0.20.3 that is characteristic of mixing associated with shear instabilities
(e.g., Linden 1979, Figure 4, where the flux Richardson number has been averaged
over a sufficiently long timescale for equality to be obtained between Rf and E c).
Particularly for flows with lower characteristic stratification, this phase of flow
evolution dominates the overall mixing character of the ultimate flow. In the next
section, we discuss this late stage in more detail, drawing connections between
the properties of this late stage with experimental and theoretical investigations
of mixing within stratified flows, paying particular attention to the implications of
the variation of mixing efficiency with stratification during this turbulent phase.

1 Nov 2002

18:38

160

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

4. TURBULENT-MIXING EFFICIENCY AS A
FUNCTION OF RICHARDSON NUMBER
Because the late, turbulent stage of flow evolution appears to be so important, it
is therefore useful to redefine the cumulative mixing efficiency as a function of
time t subsequent to the onset of true three-dimensional turbulent mixing. This
quantity, denoted Etc , is shown in Figure 10a for several flows that differ from one
another in terms of their initial Richardson number. Inspection of these results
shows that there exists clear evidence of nonmonotonicity of the mixing efficiency
as a function of the initial Richardson number. Also shown in Figure 10bd are
the mean fields of density, horizontal velocity, and B at the time t t3Dmax = 40
for the three-dimensional flows with Ri(0) = 0.025, 0.05, and 0.1, respectively,
showing clear evidence of the development of an intermediate layer of well-mixed
fluid when Ri(0) = 0.05.
Averaging over the interval for which results are shown in Figure 10a for the
cumulative mixing efficiency, we obtain the efficiency versus initial Richardson
numbered diagram shown in Figure 11a. Here we have plotted the results obtained
for seven nonzero values of the initial Richardson number. Inspection of the results
plotted on Figure 11a clearly shows that the efficiency with which the density
field is mixed in the turbulent regime is a strongly nonmonotonic function of the
stratification and moreover that the range of efficiencies that we have inferred is
in accord with the values that have been obtained from laboratory measurements.
As originally conjectured by Phillips (1972) and Posmentier (1977), nonmonotonicity in the mixing efficiency should lead to the development of distinct layers
within the stratified fluid. Physically, this can be understood (see Linden 1979 for
a clear discussion) by considering a location where the density gradient is locally
increased relative to neighboring values. If the mixing efficiency is a decreasing
function of stratification in this vicinity, the flux of buoyancy above and below this
particular level will be greater, where the stratification is weaker, thus leading to
accumulation at this location and hence intensification of the local gradient.
Such nonmonotonicity and dependence of the mixing efficiency purely on the
overall stratification requires that the dominant scales of mixing should be small
compared to the density gradient. Although this is not true of the early stages of
shear-flow development (where the characteristic scale of the flow is that of the
primary billow), the evidence from the numerical simulations points to the scales
of mixing at later times being substantially smaller than the injection scale of
the streamwise vortices characteristic of the secondary convective instability, thus
indeed being smaller than the stratification scales.
It is unfortunate that the simple physical model presented by Linden (1979)
leads to a mathematically ill-posed problem, as pointed out by Barenblatt et al.
(1993). They showed, however, that if a finite time lag existed for the turbulence to
adjust to the ambient stratification, then the problem became well posed, leading
to the development of finite layers. Consideration of the late-time behavior of
a stratified shear layer points to just such a finite time lag, as can be inferred

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

161

Figure 10 (a) Variation of the posttransition cumulative mixing efficiency (Ect ) with t-t3dmax
(listed in Figure 8) for inherently three-dimensional simulations with Ri(0) = 0.025 (solid
line), 0.05 (dashed line), and 0.1 (dot-dashed line). Vertical profiles of horizontally averaged
velocity (dotted line), horizontally averaged density (solid line), and background density B
(dot-dashed line) at the time t-t3dmax = 40 (listed in Figure 8) for three-dimensional flows
with Ri(0) = (b) 0.025, (c) 0.05, and (d) 0.1. [Adapted from figure 10 of Caulfield & Peltier
(2000), used with permission.]

from the mixing data shown in Figure 9. As already mentioned, nonzero available
potential energy is a necessary condition for irreversible mixing processes to occur
within the fluid, and such intermediate creation of P A occurs on the timescale
of the turnover of the turbulent eddies within the flow, yielding a mechanism
consistent with that postulated by Barenblatt et al. As is clearly demonstrated in
Figure 10, an intermediate well-mixed layer does indeed develop for intermediate
stratifications.
A further important point that needs to be noted is that there is a qualitative
change in the behavior of the flow for larger values of Ri(0), with the efficiency
increasing again at higher stratifications. This appears to be due to the fact that the
dissipation is decreasing owing to the reduced intensity of the streamwise vortices
at late times, whereas the quantity of mixing associated with this breakdown does

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

162

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Figure 11 (a) Variation of the time-averaged value of the posttransition cumulative mixing
efficiency (Ect ) with Richardson number for the seven simulations shown in Figure 9 (the
connecting curve is generated by cubic spline interpolation). (b) Schematic illustration of
three possible relations between the buoyancy gradient and buoyancy flux for turbulentmixing experiments as proposed by Balmforth et al. (1998). The three curves correspond to
the assumption that forcing exerts constant force per unit mass (dashed line), with constant
power (dotted line) or equipartition (solid line) implying turbulent eddies adjust to an
external velocity scale on the eddy turnover timescale. [Figure 11b is adapted from figure 1
in Balmforth et al. (1998), used with permission.]

not decrease, owing at least in part to the increased stratification and, hence, to the
increased characteristic density differences in the flow.
This behavior is somewhat suggestive of the variation of buoyancy flux with
overall stratification that was recently conjectured by Balmforth et al. (1998) in an
attempt to develop a model to explain the layer formation observed in experiments
attributed to Park et al. (1994). Balmforth et al. (1998) postulated, using mixinglength theory, that such a variation in buoyancy flux (as shown in Figure 11b) could
occur if the characteristic eddy speed for turbulent motions within a stratified fluid
adjusted to an external driving velocity (in this case, that of the stirring rods)
on the timescale of an eddy turnover. Within the framework of a stratified shear
flow, the analogous assumption is that the streamwise vortices, associated with
the convective instability, are intensified (and hence driven to mixing) by the
background shear on a timescale comparable with their turnover time, essentially
that which occurs during the relatively rapid period of turbulence onset within the
flow. This gives further support to the observation that stratified mixing driven
by the breakdown of secondary streamwise vortices may lead to the creation of a
staircase density structure.

5. CONCLUSIONS
In this paper, we review and somewhat extend our own recent work on the inference
of mixing efficiency in stratified parallel flows that undergo the process of turbulent
collapse. This process begins with the onset of a two-dimensional, essentially

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

163

inviscid, KH instability. As the two-dimensional KH billows saturate, the rate of


change with time of the two-dimensional fields is sufficiently slow to allow for
the emergence of a secondary instability. This secondary instability is convective
in nature and is initially focused in the regions of overturning isopycnals that
are induced by the rolling-up of the vorticity into the cats-eye structures that
characterize the individual billows. It also consists of streamwise vortex streaks of
alternating sign in the transverse direction. As these streamwise vortices amplify,
they are stretched in the streamwise direction and eventually interact with those
associated with the adjacent billow core. This final stage of interaction results in
the dramatic loss of spatial coherence and the sharp rise of viscous dissipation that
characterizes the mixing transition, and particularly for smaller values of the
flow stratification, this late stage of evolution dominates the mixing within the
flow.
As demonstrated through discussion of the detailed theoretical analyses in Section 3, the efficiency of the mixing process varies considerably during the three
primary stages of transition. Immediately after the initial stage of two-dimensional
KH instability amplification, the efficiency of mixing according to our definition
is extremely high, although the amount of mixing can be small, owing to the fact
that this stage is fundamentally laminar. This mixing is associated with the nucleation of the secondary instabilities, which play the essential role in the mixing
transition. During the final stage of three-dimensional turbulent collapse, on the
contrary, the mixing efficiency is small and on the order of 0.2, in accord with the
experimental measurements. In this stage, however, the amount of mixing can be
high, provided the stratification is sufficiently weak for intense development of
the secondary instabilities. The stage of flow evolution through which these two
asymptotic regimes of behavior are connected, that which is associated with the
three-dimensional coherent structures that are referred to in the experimental literature as vortex streaks, is clearly critical to the transition process. It is during this
short-lived regime that viscous dissipation first begins to rise significantly because
the spatial scale of these streaks is an order of magnitude less than the spatial scale
of the parent KH billow.
It is important to note that the efficiency with which the density field is mixed
during the mixing transition in stably stratified parallel flow is a nonmonotonic
function of the initial minimum gradient Richardson number of the flow. Both
Philips (1972) and Posmentier (1977) have proposed on theoretical grounds that
such nonmonotonicity should be a generic property of the mixing induced by
small-scale turbulence. Barenblatt et al. (1993) have specifically argued that such
dependence inevitably leads to the formation of a vertical staircase variation
of density, provided there is an adjustment time of the turbulence to the ambient
stratification, which is also characteristic of the stratified shear-flow transition
problem. However, theoretical, numerical, and experimental suggestions indicate
(see Balmforth et al. 1998, Caulfield & Kerswell 2001, Strang & Fernando 2001)
(also see Figure 11) that there is a further regime to be explored at higher ambient
stratifications and that more complicated background flows may lead to higher
efficiencies of, still turbulent, mixing.

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

164

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

Indeed, strong experimental evidence for the occurrence of a change in the


behavior of the mixing as the primary KH billows reduce in importance has been
presented recently by Strang & Fernando (2001). When their flow was visually
susceptible to primary KH instabilities, they found measured values of appropriately averaged flux Richardson numbers (which could be identified with Ect , as
shown in Figures 10 and 11) of the order of 0.20.3. However, with their particular
experimental design, increasing total stratification inevitably led to a reduction in
the relative depth of the region of density variation, and so for larger values of ambient stratification, Holmboe waves also developed. The interaction between the
two coexisting instabilities led to substantially higher mixing efficiencies, on the
order of 0.40.5 for intermediate stratifications. As the stratification was further
increased, the mixing efficiency then dropped substantially, providing evidence
that such strongly stratified flows were not greatly susceptible to flow instabilities
of any kind. Nevertheless, the accumulating evidence suggests that layering in
the stratification and efficiencies of the order of 0.2 are characteristic of mixing
catalyzed by the development of KH instabilities and are driven by the associated
secondary convective instabilities.
In this review, we illustrate the crucial variations of mixing efficiency with
stratification that occur with relatively simple initial profiles of velocity and density.
It will be necessary to consider the mixing associated with a much wider range
of initial flow conditions before really useful parameterizations of this complex
problem can be developed, with three main avenues of attack being apparent. First,
the effect of variation with Prandtl number, as initially explored by Klaassen &
Peltier (1985a) and Smyth et al. (2001), will have to be investigated in detail at
sufficiently high Reynolds number and Peclet number to have an intense turbulentmixing phase, as the existence of such a phase has been shown to be crucial to the
mixing within the flow.
Second, a wider range of flow profiles must be considered because, as shown
experimentally by Strang & Fernando (2001), interactions between instabilities of
various kinds seem to lead to qualitatively different mixing properties. An interesting, still unanswered, question to be addressed is whether the mixing efficiency
associated with secondary instabilities catalyzed by primary KH instabilities is a
generic property of stratified shear flows, or is it special to the particular initial
conditions subject to primary KH instabilities. Although parameterizations, such
as those by Mellor & Yamada (1982), do essentially make this assumption, by
restricting Rf to be at most 0.3 or so, recent experimental (Strang & Fernando
2001) and observational (Pardyjak et al. 2002) evidence suggests the possibility
of mixing efficiencies of the order of 0.40.5 in certain circumstances. In other
words, the evidence points toward the possibility that the kinetic energy of the flow
may be lost equally to viscous dissipation and irreversible mixing. Such behavior
may be characteristic of a stratified Couette flow, if the flow is able to attain the
theoretical maximum upper bound for mixing (see Caulfield & Kerswell 2001),
and was postulated on scaling grounds by Townsend (1958) for statistically steady
turbulent flow. It would be of great practical interest to be able to identify such

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING

P1: IBC

165

behavior in a time-dependent simulation of a stratified turbulent flow, analyzed


using the formalism reviewed in Section 2.
Finally, and perhaps most importantly, the interaction of turbulent mixing with
the internal gravity wave spectrum must be considered (e.g., as in Sutherland &
Peltier 1994). Internal waves lead inevitably to both temporal and spatial variability
in the underlying flow, and so the effect of time variation in the background flow on
the development of shear instabilities should be investigated. Also, internal waves
provide a mechanism by which energy and momentum can be transported away
from their generation region (Sutherland & Peltier 1995), and thus the inclusion of
gravity wave dynamics within the flow may have a qualitatively important effect
on the energy and momentum budgets. Understanding the interplay between the
wave-like and eddy-like parts of realistic flows is clearly the major remaining
challenge for developing an understanding of stratified mixing and for realizing
the full geophysical relevance of studies such as those reviewed and extended here.
The Annual Review of Fluid Mechanics is online at http://fluid.annualreviews.org

LITERATURE CITED
Arendt S, Fritts DC, Andreassen O. 1997.
The initial value problem for Kelvin vortex
waves. J. Fluid Mech. 344:181212
Balmforth NJ, Llewellyn Smith SG, Young
WR. 1998. Dynamics of interfaces and layers
in a stratified turbulent fluid. J. Fluid Mech.
355:32958
Barenblatt GI, Bertsch M, Dal Paso R, Prostokishin VM, Ughi M. 1993. A mathematical
model of turbulent heat and mass transfer in
stably stratified shear flow. J. Fluid Mech.
253:34158
Browand FK, Winant CD. 1973. Laboratory observations of shear instability in a stratified
fluid. Bound.-Layer Meteorol. 5:6777
Caulfield CP, Kerswell RR. 2001. Maximal
mixing rate in turbulent stably stratified Couette flow. Phys. Fluids 13:894900
Caulfield CP, Peltier WR. 1994. Threedimensionalization of the stratified mixing
layer. Phys. Fluids 6:38035
Caulfield CP, Peltier WR. 2000. The anatomy
of the mixing transition in homogeneous and
stratified free shear layers. J. Fluid Mech.
413:147
Cortesi AB, Yadigaroglu G, Bannerjee S. 1998.
Numerical investigation of the formation of

three-dimensional structures in stably stratified mixing layers. Phys. Fluids 10:144973


Crow SC. 1970. Stability theory for a pair of
trailing vortices. AIAA J. 8:2172
Davis PA, Peltier WR. 1979. Some characteristics of the Kelvin-Helmholtz and resonant
over-reflection modes of shear flow instability and of their interaction through vortex
pairing. J. Atmos. Sci. 36:2394412
De Silva IPD, Fernando HJS, Eaton F, Hebert
D. 1996. Evolution of Kelvin-Helmholtz billows in nature and laboratory. Earth Planet.
Sci. Lett. 143:21731
Diamessis PJ, Nomura KK. 2002. The structure
and dynamics of overturns in stably stratified homogeneous turbulence. J. Fluid Mech.
Submitted
Dimotakis PE. 2000. The mixing transition in
turbulent flows. J. Fluid Mech. 409:6998
Drazin PG, Reid WH. 1981. Hydrodynamic
Stability. Cambridge, MA: Cambridge Univ.
Press
Fernando HJS. 1991. Turbulent mixing in stratified fluids. Annu. Rev. Fluid Mech. 23:455
93
Fritts DC, Arendt S, Andreassen O. 1998. Vorticity dynamics in a breaking internal gravity

1 Nov 2002

18:38

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

166

AR

AR159-FM35-08.tex

PELTIER

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

P1: IBC

CAULFIELD

wave. Part 2. Vortex interactions and transition to turbulence. J. Fluid Mech. 367:4765
Gerz T, Schumann U, Elghobashi SE. 1989. Direct numerical simulation of stratified homogeneous turbulent shear flows. J. Fluid Mech.
200:56394
Holmboe J. 1962. On the behaviour of symmetric waves in stratified shear layers. Geofys.
Publ. Oslo 24:67113
Holt SE, Koseff JR, Ferziger JH. 1992. A numerical study of the evolution and structure
of homogeneous stably stratified sheared turbulence. J. Fluid Mech. 237:499539
Howard LN. 1961. Note on a paper of John W.
Miles. J. Fluid Mech. 10:50912
Jacobitz FG, Sarkar S, Van Atta CW. 1997. Direct numerical simulations of the turbulence
evolution in a uniformly sheared and stably
stratified flow. J. Fluid Mech. 342:23161
Keller KH, Van Atta CW. 2000. An experimental investigation of the vertical temperature
structure of homogeneous stratified shear turbulence. J. Fluid Mech. 425:129
Kerswell RR. 2002. Elliptical instability. Annu.
Rev. Fluid Mech. 34:83113
Klaassen GP, Peltier WR. 1985a. Prandtl number effects on the evolution and stability of
finite amplitude Kelvin-Helmholtz billows.
Geophys. Astrophys. Fluid Dyn. 32:2360
Klaassen GP, Peltier WR. 1985b. The onset
of turbulence in finite amplitude KelvinHelmholtz billows. J. Fluid Mech. 155:135
Klaassen GP, Peltier WR. 1989. The role of
transverse secondary instabilities in the evolution of free shear layers. J. Fluid Mech.
202:367402
Klaassen GP, Peltier WR. 1991. The influence
of stratification on secondary instabilities in
free shear layers. J. Fluid Mech. 227:71106
Koop G, Browand FK. 1979. Instability and turbulence in stratified fluids. J. Fluid Mech.
93:13559
Lawrence GA, Browand FK, Redekopp LG.
1991. The stability of a sheared density interface. Phys. Fluids A 3:236070
Linden PF. 1979. Mixing in stratified fluids.
Geophys. Astrophys. Fluid Dyn. 13:223
Lorenz EN. 1955. Available potential energy

and the maintenance of the general circulation. Tellus 7:15767


Mellor GL, Yamada T. 1982. Development of
a turbulence closure model for geophysical
fluid problems. Rev. Geophys. Space Phys.
20:85175
Miles JW. 1961. On the stability of heterogeneous shear flows. J. Fluid Mech. 10:496
508
Palmer TL, Fritts DC, Andreassen O. 1996.
Evolution and breakdown of KelvinHelmholtz billows in stratified compressible
flows. 1. Instability structure, evolution and
energetics. J. Atmos. Sci. 53:3192212
Palmer TL, Fritts DC, Andreassen O, Lie
I. 1994. Three-dimensional evolution of
Kelvin-Helmholtz billows in stratified compressible flow. Geophys. Res. Lett. 21:2287
90
Pardyjak ER, Monti P, Fernando HJS. 2002.
Flux Richardson number measurements in
stable atmospheric shear flows. J. Fluid
Mech. 459:30716
Park YG, Whitehead JA, Gnanadeskian A.
1994. Turbulent mixing in stratified fluids, layer formation and energetics. J. Fluid
Mech. 279:279312
Peltier WR, Halle J, Clark TL. 1978. The evolution of finite-amplitude Kelvin-Helmholtz
billows. Geophys. Astrophys. Fluid Dyn.
10:5387
Phillips OM. 1972. Turbulence in a strongly
stratified fluid is unstable? Deep Sea Res.
19:7981
Pierrehumbert RT, Widnall SE. 1982. The twoand three-dimensional instabilities of a spatially periodic shear layer. J. Fluid Mech.
114:5982
Posmentier ES. 1977. The generation of salinity
fine structures by vertical diffusion. J. Phys.
Oceanogr. 7:298300
Potylitsin P, Peltier WR. 1998. Stratification effects on the stability of columnar vortices on
the f-plane. J. Fluid Mech. 355:4579
Potylitsin P, Peltier WR. 1999. Three-dimensional destabilization of Stuart vortices: the
influence of rotation and ellipticity. J. Fluid
Mech. 387:20526

1 Nov 2002

18:38

AR

AR159-FM35-08.tex

AR159-FM35-08.sgm

LaTeX2e(2002/01/18)

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

STRATIFIED SHEAR FLOW MIXING


Scinocca JF. 1995. The mixing of mass and
momentum by Kelvin-Helmholtz billows. J.
Atmos. Sci. 52:250930
Smyth WD, Klaassen GP, Peltier WR. 1988.
Finite amplitude Holmboe waves. Geophys.
Astrophys. Fluid Dyn. 43:181222
Smyth WD, Moum JN. 2000a. Anisotropy of
turbulence in stably stratified mixing layers.
Phys. Fluids 12:134362
Smyth WD, Moum JN. 2000b. Length scales of
turbulence in stably stratified mixing layers.
Phys. Fluids 12:132742
Smyth WD, Moum JN, Caldwell DR. 2001.
The efficiency of mixing in turbulent patches:
inferences from direct simulations and microstructure observations. J. Phys. Oceanogr.
31:196992
Smyth WD, Peltier WR. 1994. Three-dimensionalization of barotropic vortices on the
f-plane. J. Fluid Mech. 265:2564
Smyth WD, Winters KB. 2002. Turbulence
and mixing in Holmboe waves. J. Phys.
Oceanogr. Submitted
Strang EJ, Fernando HJS. 2001. Entrainment
and mixing in stratified shear flows. J. Fluid
Mech. 428:34986
Sutherland BR, Caulfield CP, Peltier WR.
1994. Internal wave generation and hydrodynamic instability. J. Atmos. Sci. 51:3261
80
Sutherland BR, Peltier WR. 1994. Turbulence
transition and internal wave generation in

P1: IBC

167

density stratified jets. Phys. Fluids A 6:1267


84
Sutherland BR, Peltier WR. 1995. Internal
wave emission into the middle atmosphere
from a model tropospheric jet. J. Atmos. Sci.
52:321435
Thorpe SA. 1977. Turbulence and mixing in a
Scottish loch. Philos. Trans. R. Soc. London
Ser. A 286:12581
Thorpe SA. 1987. Transitional phenomena and
the development of turbulence in stratified
fluids: a review. J. Geophys. Res. 92C:5231
48
Townsend AA. 1958. The effects of radiative
transfer on turbulent flow of a stratified fluid.
J. Fluid Mech. 4:36175
Tseng Y-H. Ferziger JH. 2001. Mixing and
available potential energy in stratified flows.
Phys. Fluids 13:128193
Werne J, Fritts DC. 1999. Stratified shear turbulence: evolution and statistics. Geophys. Res.
Lett. 26:43942
Winters KB, DAsaro EA. 1996. Diascalar flux
and the rate of fluid mixing. J. Fluid Mech.
317:17993
Winters KB, Lombard PN, Riley JJ, DAsaro
EA. 1995. Available potential energy and
mixing in density-stratified fluids. J. Fluid
Mech. 289:11528
Woods JD. 1968. Wave-induced shear instability in the summer thermocline. J. Fluid Mech.
32:791800

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

29 Nov 2002

13:41

AR

AR159-08-COLOR.tex

AR159-08-COLOR.SGM

LaTeX2e(2002/01/18)

P1: GDL

Figure 3 Isosurfaces of spanwise vorticity (green), streamwise vorticity of opposite


signs (red and blue), and absolute value of vertical vorticity (yellow) for a simulation
with Re = 750, Pr = 1, and Ri(0) = 0.05 at three characteristic times (t2dmax): when
the two-dimensional perturbation has maximal amplitude, when the kinetic energy
associated with the three-dimensional perturbation is 1% of the kinetic energy associated with the two-dimensional perturbation, and when the dissipation within the flow
is maximal.

13:41

AR

AR159-08-COLOR.tex

AR159-08-COLOR.SGM

LaTeX2e(2002/01/18)

P1: GDL

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

29 Nov 2002

Figure 6 Isosurfaces of spanwise vorticity (green) and positive (blue) and negative
(red) streamwise vorticity within (a) unstratified and (b) stratified (with Ri(0) = 0.05)
simulations containing two neighboring primary Kelvin-Helmholtz billows, during the
development of intense spanwise-periodic streamwise vortices.

P1: FDS

November 22, 2002

11:16

Annual Reviews

AR159-FM

Annual Review of Fluid Mechanics


Volume 35, 2003

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

CONTENTS
STANLEY CORRSIN: 19201986, John L. Lumley and Stephen H. Davis
AIRCRAFT ICING, Tuncer Cebeci and Fassi Kafyeke
WATER-WAVE IMPACT ON WALLS, D. H. Peregrine
MECHANISMS ON TRANSVERSE MOTIONS IN TURBULENT WALL
FLOWS, G. E. Karniadakis and Kwing-So Choi
INSTABILITIES IN FLUIDIZED BEDS, Sankaran Sundaresan
AERODYNAMICS OF SMALL VEHICLES, Thomas J. Mueller
and James D. DeLaurier

MATERIAL INSTABILITY IN COMPLEX FLUIDS, J. D. Goddard


MIXING EFFICIENCY IN STRATIFIED SHEAR FLOWS, W. R. Peltier
and C. P. Caulfield

THE FLOW OF HUMAN CROWDS, Roger L. Hughes


PARTICLE-TURBULENCE INTERACTIONS IN ATMOSPHERIC CLOUDS,
Raymond A. Shaw

LOW-DIMENSIONAL MODELING AND NUMERICAL SIMULATION


OF TRANSITION IN SIMPLE SHEAR FLOWS, Dietmar Rempfer
RAPID GRANULAR FLOWS, Isaac Goldhirsch
BIFURCATING AND BLOOMING JETS, W. C. Reynolds, D. E. Parekh,
P. J. D. Juvet, and M. J. D. Lee

1
11
23
45
63
89
113
135
169
183
229
267
295

TEXTBOOK MULTIGRID EFFICIENCY FOR FLUID SIMULATIONS,


James L. Thomas, Boris Diskin, and Achi Brandt

317

LEVEL SET METHODS FOR FLUID INTERFACES, J. A. Sethian


and Peter Smereka

341

SMALL-SCALE HYDRODYNAMICS IN LAKES, Alfred Wuest


and Andreas Lorke

STABILITY AND TRANSITION OF THREE-DIMENSIONAL BOUNDARY


LAYERS, William S. Saric, Helen L. Reed, Edward B. White
SHELL MODELS OF ENERGY CASCADE IN TURBULENCE, Luca Biferale
FLOW AND DISPERSION IN URBAN AREAS, R. E. Britter and S. R. Hanna

373
413
441
469

vii

P1: FDS

November 22, 2002

viii

11:16

Annual Reviews

AR159-FM

CONTENTS

INDEXES
Subject Index
Cumulative Index of Contributing Authors, Volumes 135
Cumulative Index of Chapter Titles, Volumes 135

ERRATA

Annu. Rev. Fluid. Mech. 2003.35:135-167. Downloaded from arjournals.annualreviews.org


by University of Toronto on 06/14/06. For personal use only.

An online log of corrections to Annual Review of Fluid Mechanics chapters


may be found at http://fluid.annualreviews.org/errata.shtml

497
521
528

Anda mungkin juga menyukai